Вы находитесь на странице: 1из 22

1.

Fundamentals of grain boundaries and grain boundary migration


1.1. Introduction

The properties of crystalline metallic materials are determined by their


deviation from a perfect crystal lattice, which occurs due to their intrinsic crystal
defects. Without those crystal defects, metals would never have been used in
such a wide variety of applications [2]. There are different types of crystal
defects which are distinguished by their spatial dimension. One of the longest
known and most important defects in metals is the grain boundary, a twodimensional planar defect, which separates two adjacent crystallites of the same
crystal structure and chemical composition, but of different orientation [1].
Since any crystalline material, except for single crystals, is granular-structured,
grain boundaries are the fundamental defect in polycrystalline materials,
exerting a substantial impact on their properties [2].

1.2. Classification
In three-dimensional space, there are eight degrees of freedom, which means
eight independent parameters are required to assign a mathematically exact
definition to a given grain boundary [9]. Three parameters are necessary to
define the orientation relationship between two adjacent grains which are
usually represented by an Euler angle triplet (M1, ), M2), by Miller indices or by
an angle-axis pair in Rodrigues-Frank space [2, 3, 4].
Two more parameters are needed to define the spatial orientation of the grain
boundary plane, i.e. the boundary inclination, which are expressed by the
normalized crystallographic normal vector of the plane of inclination n = (n1, n2,

4
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

n3) with respect to one of the adjacent grains. Additional to these five
macroscopic parameters there are three independent values of the microscopic
translational vector t = (t1, t2, t3) [2]. All intrinsic properties of grain boundaries,
particularly mobility and energy, are functions of these eight parameters [1]. The
five macroscopic parameters may be influenced externally, whereas the
translational vector t is determined by the minimum of the total energy of the
crystal.
To determine the dependence of grain boundary properties on the five
macroscopic parameters introduced above, it would be necessary to keep all but
one degree of freedom constant and successively vary one free parameter [1, 2].
However, under realistic experimental conditions only one set of parameters
may remain at a fixed value. The most common investigation procedure
therefore is to change an orientation relationship by varying the rotation angle
and keeping the plane of inclination fixed with respect to a reference point or
vice versa while preserving one common axis of rotation [1].

1.3. Presentation of misorientations


The misorientation or orientation relationship between two differently
oriented crystallites is a spatial transformation applied to one crystal in order to
bring both orientations to coincidence [1]. Assuming a common origin for both
lattices, this is achieved by a simple rotation of one lattice relative to the other,
which is conveniently expressed by a rotation transformation matrix g m :

5
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

^Ci `

^ `

gm Cj

(1.1)

^ `

where ^Ci ` and C j are the crystal coordinate systems. The easiest way to
describe such a rotation matrix is in terms of a rotation axis <hkl> common for
both lattice coordinate systems and a rotation angle T (Fig. 1-1), since for many

z z

 uvtw !

x
y
x
Fig. 1-1:

Rotation about axis <uvtw> of angle T


brings coordinate system 1 and 2 to coincide.

cases it is of major interest to investigate the influence of the rotation angle on a


grain boundary property for a certain rotation axis [1]. Thus, it is common to
keep the grain boundary inclination constant and only take into account the
rotation angle dependence. Consequently, grain boundaries may be divided into
three different categories. In case of a rotation axis perpendicular to the grain
boundary plane, the grain boundary is given the denotation twist boundary (Fig.
1-2 a). For this type of grain boundary the grain boundary plane is exactly
defined and independent of the rotation angle. In Figures 1-2 b and 1-2 c the
class of tilt grain boundaries is depicted. The major difference to twist

6
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

boundaries is that the rotation axis is aligned parallel to the boundary plane [1].
Consequently, an infinite number of possible grain boundary planes exists for a
given rotation angle. Further, one distinguishes between two types of grain
boundary within the class of tilt boundaries. Under the special conditions that
mirror symmetry is satisfied between two adjacent grains, the boundary is
labeled as a symmetrical tilt boundary (Fig. 1-2 c). All other configurations are
denoted as asymmetrical tilt boundaries (Fig. 1-2 b) [1].

rotation axis

(a)

grain
boundary

(b)
symmetry plane

[uvtw]1

[uvtw]2

grain
boundary
tilt axis and

(c)

boundary

plane

grain boundary
rotation axis

Fig. 1-2:

Schematic depictions of a) a twist grain boundary and b) an


asymmetrical tilt and c) a symmetrical tilt boundary [1]

Grain boundaries that do not satisfy either of the criteria of the first two
classes are termed mixed or random grain boundaries and consist of both twist
and tilt components.

7
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

1.4. Atomic structure


An additional means to classify grain boundaries is to categorize them,
depending on their misorientation, by distinguishing between low angle grain
boundaries (LAGBs) and high angle boundaries (HAGBs). In the following
sections the difference between these grain boundaries regarding their
microstructural configuration will be addressed.

1.4.1. Low angle grain boundaries


Providing the case that the misorientation angle between two neighboring
grains is small enough (LAGB), the misorientation of the grain boundary is
achieved by lattice distortions introduced by dislocation arrays. In case of a
symmetrical  10 1 0 ! tilt boundary, the configuration consists of a single set of
edge dislocations with Burgers vectors b , where the dislocation spacing directly
correlates with the misorientation angle T (Fig. 1-3a) [1].

b
d

2 sin

T
2

(2.1)

Accompanied with an increasing misorientation angle T the spacing d


decreases as depicted in Fig. 1-3.

8
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

boundary

bA
d = b/MT

grain 1

grain 2

dislocation spacing d [10-4 cm]

6
5
4
3

measured
. .10 -8
40
d= M

2
1

M
T
0
0
(a)

(b)

2
3
4
4
tilt angle [10 rad]

Fig. 1-3: a) Schematic depiction of a symmetrical low angle tilt boundary, b)


measured and calculated dislocation spacing versus tilt angle T in a
symmetrical LAGB in Germanium [1]

For asymmetrical tilt grain boundaries (Fig. 1-2 b), in which the boundary
plane deviates from its symmetrical position by an inclination angle M (Fig. 14), at least two sets of edge dislocations are necessary to accommodate the
boundary configuration. The Burgers vector of these two dislocation sets must
be perpendicular to each other (Fig. 1-4 a) and with increasing asymmetry, the
fraction of the second set of dislocations ( b2 ) has to increase,

1
d2

b2

T sin M

(2.2)

while the fraction of dislocations with Burgers vector b1 decreases with M .

9
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

a)

b)

 uvtw !1

M  1 / 2T

T/2

plane of symmetry
T/2

M
M  1 / 2T

 uvtw ! 2

Fig. 1-4:

boundary plane

a) Lattice dislocation arrangement [1] and b) rotation angle T and


inclination angle M of an asymmetrical tilt LAGB.

b1
1
d1 T cosM

(2.3)

Small angle twist boundaries require two sets of screw dislocations and in the
most general case of mixed LAGBs the boundary structure is comprised of
dislocation networks of three Burgers vectors [5]. By applying this dislocation
model of LAGBs, which was developed by Read and Shockley [6], the exact
calculation of free grain boundary energy is possible. As derived by Read and
Shockley, the stress field of dislocations in an infinite periodic array is spatially
limited to a range of the order of spacing d . In case of an edge dislocation its
energy Ed per unit length is thus expressed by:

10
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

Ed

Pb 2
d
ln  Ec
4S 1 Q r0

(2.4)

where Q is the Poisson ratio, P the shear modulus, r0 | b the radius and Ec the
energy of the dislocation core [1]. Keeping in mind, that this equation needs to
be applied to a set of dislocations, which compose a LAGB, we introduce the
number of dislocations per unit length n { 1 / d

T / b to the above equation

and thereby obtain the energy per unit area for the case of a symmetrical tilt
grain boundary:

J bsymm

where A

1
T Pb 2

ln
E

c T A  B ln T
b 4S (1 Q ) T

(2.5)

Ec / b and B Pb / 4S (1 Q ) . Comparing the energies obtained by

calculation with equation 2.4 to experimental data [7] proves the validity of this
dislocation model, as the energy increase with increasing angle of rotation T is
predicted, accordingly, for angles T  15q [1].

1.4.2. High angle grain boundaries


For rotation angles larger than 15, the lattice dislocation model no longer
applies, as the dislocation cores are brought to an overlapping [8], causing the
loss of their identity as individual lattice defects. Therefore, grain boundaries
with rotation angles larger than 15 are distinguished from LAGBs and are

11
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

termed high angle grain boundaries (HAGB). The current concept of the
structure of HAGBs was deduced from geometrical considerations, based on
dislocation models for LAGBs and atomistic simulations [5, 7, 9, 1].
Most commonly, the structure of interfaces is described by the coincidence
site lattice (CSL)- displacement shift complete (DSC) lattice construction. The
atomic arrangement in a perfect crystal is determined by the minimum of free
energy and any deviation of the former from ideal positions inevitably causes an
increase of the latter [2]. Consequently, it is safe to assume that the atoms will
remain close to their ideal position. At certain misorientations between two
crystallites, crystallographic planes exist that transcend the grain boundary from
one crystallite to the other, i. e. certain atomic positions in the grain boundary
coincide with ideal positions in both neighboring crystallites. These atomic
positions are termed coincidence sites and the super lattice containing these sites
on the other hand is the aforementioned coincidence site lattice (CSL). The
elementary cell of the CSL is self-evidently larger than the elementary cell of
the crystal lattice and its volume may be calculated in terms of the lattice
parameters of the crystal lattice. The CSL is characterized by the density of its
coincidence sites, which in turn are defined by the quantity 6 [2]:

volume of elementary cell CSL


volume of elementary cell of crystal lattice

(2.6)

Since coincidence sites are representations of atoms located at ideal fit


positions, it is safe to assume that grain boundaries preferably stretch along
coincidence sites [2].

12
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

a
a

a 5a 5

DSC latticeDSC

CSL CSL

CSL CSL

a
a
a

lattice point grain I

a)
Fig. 1-5:

lattice point grain II

b)

a) Coincidence site lattice (CSL) of a 36.87 <100> grain boundary


=5). Left: grain boundary plane is perpendicular to paper plane (tilt
(
boundary). Right: grain boundary plane is parallel to paper plane
(twist boundary). b) CSL and displacement shift complete lattice (DSC)
at 36.87 <100> rotation in a cubic lattice. [1]

Grain boundaries containing a high density of coincidence sites are called


CSL boundaries or special boundaries. The order of the boundary increases with
the decrease of the 6value, which is always an odd integer. Experimental data
show that CSL grain boundaries of low 6indeed consist of a low energy
configuration, which is reflected by a low free surface energy [10, 11] and low
grain boundary migration activation enthalpy [12]. However, there is a
fundamental problem in applying this concept to real, arbitrary grain boundaries:
since the CSL only exists at very special, defined angles and does not
continually change with misorientation angle T, the long range coincidence is
lost even for small deviations [2]. This may be compensated by the introduction
of dislocations with DSC Burgers vectors which are also referred to as

13
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

secondary grain boundary dislocations (SGBDs), as they are confined to the


grain boundary [1].
Detailed analysis revealed that 7 different polyhedra (see. Fig. 1-6) are
necessary to account for all possible arrangements of atoms in a grain boundary
[13].

(D)

(H)
(E)
(])

(J)

Fig. 1-6:

(G)

(K)

The seven Bernal structures, of which the structural units of GBs are
composed [13, 38].

These polyhedra represent characteristic structures of the grain boundary and


are thus termed structural units. It has been shown by computer simulations, that
low energy boundaries consist of only one type of structural unit [2]. Upon
changing the orientation relationship, other structural units are introduced into
the boundary plane, which are identical to the cores of SGBDs. The density of
the newly introduced structural units increases with the grain boundaries
misorientation angle, until eventually the grain boundary structure solely

14
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

consists of these new structural units. This concept of structural units constitutes
our current understanding of grain boundary structures [1].
Analogously to the dislocation model of LAGBs, where the model by Read
and Shockley [6] applies, the CSL-DSC model is no longer valid above a
maximum deviation 'Z from exact CSL-misorientations, because the spacing of
the SGBDs decreases with increasing deviation. As such, the SGBDs above a
certain deviation of the misorientation angle tend to overlap and therefore lose
their individuality. The most common criterion for the maximum deviation
proposed by Brandon [15] reads:

'Z

15q
6

(2-7)

Grain boundaries satisfying this criterion are referred to as special grain


boundaries as opposed to random grain boundaries.

1.5. Classical model of grain boundary motion


Over the past 60 years, grain boundary motion has been subjected to
extensive experimental and theoretical investigations. However, to this day no
unified theory about grain boundary migration has been published, which were
able to account for phenomena such as the misorientation dependence of grain
boundary mobility as observed for instance by Aust and Rutter [16]. Basically,
all theoretical attempts to describe grain boundary motion are based on the
reaction rate theory by Smoluchoski [17] and Turnbull et al. [18], in which
individual atoms cross the grain boundary accompanied by a net energy gain. It
15
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

is assumed that the detachment of one atom from a crystal to join the crystal
across the boundary destroys a lattice site rather than creating a vacancy and that
its attachment to the adjacent crystal surface generates a new lattice site as
opposed to eliminating a vacancy [1].

Grain boundary

Grain 1

free energy G

Grain 2
Gm
Gm

p 3
Pb

Fig. 1-7: The free energy G of a moving atom


changes by the driving force pb3 when it
crosses the grain boundary. Gm is the free
energy barrier for bulk diffusion [1].

Following these assumptions, grain boundary motion is conveniently


simplified to diffusive motion of single atoms across the boundary rather than
the uniform motion of atom groups.
We shall now derive the rate equation for grain boundary velocity, as seen in
Ref. [1]: for the simplest case, the grain boundary is assumed to have a thickness
of a monatomic layer, hence the boundary may be crossed by a single atomic
jump and each transferred atom displaces the boundary by the diameter of an
atom. Grain boundary velocity v is then expressed by [19]

16
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

v b   

(2.8)

where *+ and *- are the jump frequencies in opposite directions, respectively. In


case of zero difference in Gibbs free energy between the two grains, the net flux
will also equal zero. However, any difference in Gibbs free energy per unit
volume gives rise to a driving force p:

dG
dV

(2.9)

In that case, each atom attaching to the growing grain of volume :3 | b3 will
gain the energy pb3 The change in Gibbs free energy associated with this
process is schematically shown in Fig. 1-7. The grain boundary velocity then
reads:

Gm  pb3
Gm

  exp 
v b  exp 
kT

kT

(2.10)

The assumption of equal jump frequencies Q+ = Q- = QD (QD Debye-frequency)


and equal migration free energy Gm for both jump directions simplifies the
equation toG

pb 3
G

bQ D exp  m 1  exp 
kT
kT

(2.11)

17
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

Further assuming that pb3<<kT for T t 0.3Tm , a series expansion of the second
exponential term

pb 3
pb 3
# 1 
exp 
kT
kT

(2.12)

finally yields for velocity v

b 4 vD
G
v
exp  m p { m p
kT
kT

(2.13)

where m is referred to as grain boundary mobility, the proportionality factor of


the fundamental linear relationship between grain boundary velocity and driving
force. Variable m basically contains all the kinetic characteristics of the grain
boundary. Furthermore, it is also distinguished between the intrinsic mobility
determined by grain boundary misorientation and inclination and the extrinsic
mobility primarily governed by impurity content and dislocation densities.
By introducing the Nernst-Einstein relation for general diffusive processes
(volume diffusion, grain boundary diffusion, etc.) without specification into the
equation, the relation between grain boundary mobility and diffusion of atoms
across the boundary reads:

b 2 Dm
m
kT

H
m0 exp 
kT

(2.14)

18
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

Dm is the diffusion coefficient for diffusion jumps through the boundary, m0

and H are the pre-exponential factor and activation enthalpy of the grain
boundary mobility. This simple model may be modified by assuming the
detachments to occur in a sequence of steps or thermal grain boundary vacancies
to assist diffusion [20]. However, these modifications will only affect the preexponential factor m0 and the activation enthalpy H of mobility m leaving the
migration rate v and the driving force p unaffected [1].

1.6. Driving forces of grain boundary migration


The driving force p is a force acting per unit area on a grain boundary and its
source may be of various origins. Generally, a driving force emerges when the
motion of a grain boundary results in reducing the systems overall free energy.
A gradient of any intensive thermodynamic variable offers a source of such
driving forces: temperature gradients, pressure, density of defects, energy
density (elastic, magnetic, electrical), etc. A more detailed treatise on the various
driving forces utilized in grain boundary migration experiments was written by
Gottstein and Shvindlerman [1].

1.6.1. Magnetic driving force


We shall now take a detailed look at the magnetic driving force, which was
predominantly applied to induce grain boundary migration in the experiments
featured in this work. Mullins was the first to use the driving force induced by a

19
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

high magnetic field to induce grain boundary migration in bismuth polycrystals


[21].
Due to its anisotropic properties, zinc possesses a different magnetic
susceptibility parallel ( FII ) and perpendicular ( F A ) to its hexagonal or c-axis
such that | FII |>| F A |. The susceptibility tensor is represented by:

FA

0
0

0
F||

FA
0

(2.15)

As shown schematically in Fig. 1-8 below, the susceptibility tensor surface


for zinc has an ellipsoidal shape.
Therefore, inside a magnetic field a driving force for grain boundary motion
is induced between two crystallites due to a difference of magnetic free energy
density , which is given by

P0 H  M dH

H
0

&

&

&

H2
P0
1  F ij li l j
2

(2.16)

&

where P 0 is the vacuum permeability and H is the magnetic field strength. The
&

magnetic polarization specific to the material is given by M

&

F H

(see eq.

2.15) and ls = cos4s are the cosines between the principal crystal axes and the
&

magnetic field H .

20
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

&
H

F
FH

a3

a2

a1

Fig. 1-8: Schematic depiction of the anisotropy of


magnetic susceptibility in zinc, with respect
to the unit cell. The spatial dependence of the
magnetic susceptibility may be described by
a second rank tensor and is represented by
an ellipsoidal surface [22].

As shown analogously in Ref. [22] for other anisotropic crystal properties, we


shall in the following develop a general expression for the magnetic driving
force. Taking into account, that there are only two independent susceptibility
components in zinc, equation 2.16 reduces to:

P0

H2
1  F l12  l22  F||l32
2

(2.17)

which since
l12  l22  l32 1 l12  l22 1  l32

(2.18)

may be further simplified to:

21
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

2
H2
P0
F|| - F cos 3
1  F 
2

'F

(2.19)

Thus, a general expression is obtained for the magnetic free energy density Z
associated with a given orientation of a zinc crystal in an external magnetic
field, dependent on only one spatial variable, 43, the angle between the
&

crystallographic c-axis and magnetic field vector H . To calculate the driving


force for magnetically induced grain boundary motion p, the magnetic free
energy density associated with the two orientations of the crystallites in the
magnetic field separated by the grain boundary have to be subtracted from one
another, which finally results in:

p Z1  Z2

H2
P0
'F cos2 1  cos2 2
2

(2.20)

where 41 and 4 2 are the angles between the c-axes in both neighboring
&

crystallites and the magnetic field H . The susceptibility difference in Zn was


measured to be 'F= 0.510-5 and found not to depend on the temperature [23].
Now that the driving force for bicrystalline samples was deduced, we will
proceed to magnetic driving forces in polycrystals: as applied in [24, 25], the
magnetic driving force for the growth of each grain in a polycrystal is given by
equation 2.21 below. Like discussed above for the case of bicrystalline samples,
the crystallites with energetically favorable c-axis orientation with respect to the
field are bound to grow while the crystallite with a less favorable orientation

22
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

will shrink. This is caused by the differing magnetic susceptibility in zinc


parallel and perpendicular to its crystallographic c-axis. Equation 2.21, as
introduced in Ref. [24], describes the magnetic driving force acting on one grain
in a polycrystal,

pm

Z Z

P0 H 2
2

1n
F  F j
n j

cos2 j

1
j
P0 'FH 2 cos2 
(2.21)
2
n

where 4 and 4 j are the angles between field direction and the principal axes of
the grain considered and its n neighboring grains. If 'F ! 0 (which is true for
Zn) and the condition cos2  1/ n n cos2 n  0 is met by the one grain
and its adjacent grains, the magnetic energy density Z in the one grain is lower
than the average energy density of its neighboring grains ( Z  Z ). Thus, the
magnetic driving force promotes growth of the former grain, as then pm  0
[24].
The largest obtainable driving force is governed by the magnetic field
strength at disposal and the magnetic susceptibility difference 'F of the
material under investigation. In the case of zinc, the maximum driving force pm
is calculated to 1.2 kJ/m3 in high field magnets with a maximum field strength
of 25 Tesla. Considering the applicability for grain boundary migration
experiments, two major advantages over the capillary driving force (see section
1.6.2 below) are:

23
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

1.) The magnetic driving force facilitates the addressing of absolute grain
boundary mobility as opposed to reduced mobility, thus enabling the
migration measurements of crystallographically defined grain boundaries
and studying the inclination dependence of activation parameters.
2.) The possibility to adjust the magnitude of the driving force during the
experiment.

Curvature driven migration experiments do not yield the absolute mobility as a


result, as the grain boundary inclination changes along the curvature and with it,
consequently, its structure.
Therefore, in curvature driven migration experiments one basically obtains
the averaged mobility of a large set of differently inclined grain boundaries,
rather than the absolute mobility of one spatially defined, planar boundary.

1.6.2. Capillary driving force


Unlike with grain growth induced by a magnetic driving force, where the free
energy gain during grain boundary motion originates from the alignment of
crystal volume swept by the boundary to an orientation associated with lower
magnetic susceptibility, the capillary driving force is provided by the free energy
of the grain boundary and stems from the pressure difference 'p , caused by the
difference of capillary forces on both sides of a curved grain boundary [1]. The
capillary pressure is expressed by:

24
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

1 1
'p J 
R1 R2

(2.22)

where J is the grain boundary free energy and R1 , R2 are the main radii of
curvature under the assumption of boundary isotropy. Upon thermal activation
this driving force induces boundary motion in the direction towards the center of
curvature, causing a reduction of grain boundary area and, consequently, energy.
However, the use of the capillary driving force only permits to evaluate a
reduced boundary mobility, A

Jm , where the grain boundary free energy is

still part of the equation and thus factors into the obtained mobility values.

25
'LHVHV:HUNLVWFRS\ULJKWJHVFKW]WXQGGDUILQNHLQHU)RUPYHUYLHOIlOWLJWZHUGHQQRFKDQ'ULWWHZHLWHUJHJHEHQZHUGHQ
(VJLOWQXUIUGHQSHUV|QOLFKHQ*HEUDXFK

Вам также может понравиться